A Design Based on Charge-transfer Bilayer as Electron Transport

‡Department of Electrical Engineering, National United University, Miaoli 36063, ... §College of Engineering, Chang Gung University, Taoyuan 33302,...
0 downloads 0 Views 5MB Size
Article pubs.acs.org/JPCC

Cite This: J. Phys. Chem. C 2018, 122, 236−244

A Design Based on a Charge-Transfer Bilayer as an Electron Transport Layer for Improving the Performance and Stability in Planar Perovskite Solar Cells Shang-Hsuan Wu,† Ming-Yi Lin,‡ Sheng-Hao Chang,¶ Wei-Chen Tu,¶ Chih-Wei Chu,*,†,§ and Yia-Chung Chang*,† †

Research Center for Applied Sciences, Academia Sinica, Taipei 11529, Taiwan Department of Electrical Engineering, National United University, Miaoli 36063, Taiwan ¶ Department of Electronic Engineering, Chung Yuan Christian University, Taoyuan 32023, Taiwan § College of Engineering, Chang Gung University, Taoyuan 33302, Taiwan ‡

S Supporting Information *

ABSTRACT: A highly efficient electron transport layer (ETL) is an essential constituent for good performance and stability in planar perovskite solar cells. Among n-type metal oxide materials, zinc oxide (ZnO) is a promising candidate for an electron transport layer due to its relatively high electron mobility, high transparency, and versatile nanostructures. However, it was found that several disadvantages could occur at the ZnO/perovskite interface, such as decomposition of CH3NH3PbI3 and poorly aligned energy levels. To overcome these issues, we present a design based on staircase band alignment of a low-temperature solution-processed ZnO/Al-doped ZnO (AZO) bilayer thin film as electron transport layers in planar perovskite solar cells. Experimental results revealed that the power conversion efficiency (PCE) of perovskite solar cells was significantly increased from 12.3% to 16.1% by employing the AZO thin film as the buffer layer. Meanwhile, the short-circuit current density (Jsc), open-circuit voltage (Voc), and fill factor (FF) were improved to 20.6 mA/cm2, 1.09 V, and 71.6%, respectively. The enhancement in performance is attributed to the modified interface in the ETL with staircase band alignment of ZnO/AZO/CH3NH3PbI3, which allows more efficient extraction of photogenerated electrons in the CH3NH3PbI3 active layer. Our studies demonstrated that the solution-processed ZnO/AZO bilayer ETLs provide a promising new approach for the development of low-cost, high-performance, and stable planar perovskite solar cells.



INTRODUCTION

structure. In general, the normal (n−i−p) heterojunction structure is composed of an electron transport layer−active layer−hole transport layer while in the inverted (p−i−n) heterojunction structure, the sequence of layers is reversed.24 On the other hand, a normal structure is commonly found in mesoporous heterojunctions. Among the n-type metal oxide materials used in normal perovskite solar cells, titanium dioxide (TiO2) is commonly employed in the mesoporous scaffold.25 Although TiO2-based perovskite solar cells may pave the way to high PCE devices, several disadvantages of TiO2 were reported such as low electron mobility and high annealing temperature (above 450 °C) to form a crystalline TiO 2 film in anatasephase.26 Accordingly, zinc oxide (ZnO) has been extensively studied as a substitute for TiO2 due to its similar electron affinity, relatively high electron mobility, high transparency, and versatile nanostructures.27−29

Organometal halide perovskite solar cells (e.g., CH3NH3PbX3, X = Cl, Br, I) have drawn much attention in current renewable solar research owing to their excellent optical and electronic properties, including a strong absorption band that spans the visible region,1 direct band gap (∼1.5 eV),2 long carrier diffusion length (100−1000 nm),3 high charge-carrier mobility (∼10 cm2 V−1 s−1),4 and significantly low-cost fabrication process.5,6 These superior optoelectronic properties enable an increase in power conversion efficiencies (PCEs) of planar heterojunction (PHJ) perovskite devices from 3.8% in 20097 to 22.1% in 2017.8 Due to the potential of organometal halide perovskites, there have been a great number of studies regarding the carrier dynamics of perovskites,9 development of new materials,10,11 optimization of perovskite absorber layer thickness, crystallinity, and surface coverage,12−15 and lowtemperature processing16,17 over the past few years. State-ofthe-art perovskite solar cells are based on two different device architectures: mesoporous18 and planar heterojunctions.19−23 Planar heterojunctions can either have a normal or an inverted © 2017 American Chemical Society

Received: November 14, 2017 Revised: December 10, 2017 Published: December 11, 2017 236

DOI: 10.1021/acs.jpcc.7b11245 J. Phys. Chem. C 2018, 122, 236−244

Article

The Journal of Physical Chemistry C

passivation ability of AZO,43 combined with the good transport characteristics of the ZnO ETL,46 we adopted a synergistic approach of using a low-temperature solution-processed ZnO/ AZO bilayer as an ETL which can offer better electron extraction and transportation due to the staircase band alignment and strong charge-transfer-induced field in the AZO/perovskite interface. In addition, the ZnO/AZO bilayer configuration can improve interfacial contact with the perovskite active layer and suppress charge recombination, ultimately resulting in enhancement in PCE, Jsc, and Voc of the device. Detailed investigations on the prominent effect of the ZnO/ AZO bilayer as an ETL for PHJ perovskite solar cells are presented. The Voc and Jsc values of the devices containing the ZnO/AZO bilayer were enhanced compared to those of devices containing only ZnO. Moreover, the transport properties of the ZnO and interfacial AZO layer were systematically investigated by using ultraviolet photoelectron spectroscopy (UPS) and absorption measurements. Better charge transfer was found between the ZnO/AZO bilayer and perovskite which leads to higher Voc and Jsc in the device. A maximum PCE of 16.1% was obtained for planar perovskite solar cells comprising a ZnO/ AZO bilayer ETL that is 24% higher than those containing only a ZnO ETL.

For high-performance PHJ perovskite solar cells, the selection of an electron collection layer with hole blocking capability and a low resistivity pathway for efficient electron extraction is necessary.30,31 ZnO has been demonstrated to be an effective electron collection material due to its various nanostructures that can be easily achieved by a solution process for more efficient charge extraction and transport.32,33 Recently, several studies of mesostructured perovskite solar cells based on ZnO nanorod arrays have been used to replace the mesoporous TiO2 nanostructures in conventional perovskite solar cells.34,35 For instance, Mahmood et al. have obtained a remarkable PCE of 16.1% in ZnO mesoscopic perovskite solar cells based on synergistically combining mesoscale control with nitrogen doping and surface modification of ZnO nanorod arrays.36 However, it was found that severe decomposition of CH3NH3PbI3 could occur at the ZnO/perovskite interface due to excessive OH− groups and chemical residues on the ZnO surface.37 To overcome the unstable decomposition process of CH3NH3PbI3 to PbI2 which tends to occur at the ZnO/perovskite interface, several approaches have been developed to solve the decomposition issue: (i) high temperature annealing of the ZnO layer,38 (ii) employing polymer interlayers between the ZnO/perovskite interface to avoid direct interaction,39−41 and (iii) the use of aluminum (Al) as a dopant in ZnO to passivate the OH− group.42,43 ZnO is an intrinsic n-type wide bandgap semiconductor, in which native n-type doping is due to oxygen vacancies or zinc interstitials. For the purpose of improving the electrical conductivity and optical transmittance of ZnO thin films, group III elements such as B, Al, Ga, and In are commonly substituted for Zn.44 Previously, Tseng et al. have achieved a PCE of 17.6%, which was considered to be the highest for ZnO-based perovskite solar cells by using a sputtered Al-doped ZnO (AZO) thin film as the electron transport layer (ETL).43 In the above research, it is worth noting that the use of AZO as an ETL can synergistically produce higher efficiency and stability for perovskite solar cells. However, to fabricate better thin film quality of AZO ETLs with high conductivity and crystallinity, preparation in a physical vapor deposition (PVD) system with the use of a vacuum chamber is required which can lead to an increase in production cost and hinder the scale-up technologies in roll-to-roll fabrication of large area solar cells. Therefore, the development of solution-processed AZO ETLs by sol−gel technique is desired because it presents several advantages, including facile synthesis, solution-based growth, long-term stability, and low-temperature processability that are suitable for flexible photovoltaics. To date, it has been proven that PHJ perovskite solar cells based on a TiO2 ETL show higher efficiency (>20%)45 compared to those based on a single AZO ETL (∼17%). From our point of view, using a single AZO layer as an ETL would degrade the electron transport in comparison to ZnO, because AZO is more heavily doped than ZnO. However, a lightly doped ZnO will have a high possibility for charge recombination at the ZnO/perovskite interface because of the lack of a strong built-in electric field at the interface to separate photocreated electron−hole pairs. On the other hand, by inserting a thin layer of AZO between ZnO and perovskite, a strong band bending (with high built-in electric field) near the AZO/perovskite interface occurs as a result of charge transfer from the heavy n-doped AZO to ZnO, which can greatly enhance the electron−hole separation at the interface. Inspired by this concept for effective electron extraction and the great



EXPERIMENTAL SECTION Syntheses of Materials. Methylammonium iodide (MAI) preparation in aqueous HI (57 wt % in water), methylamine (CH3NH2, 40 wt % in aqueous solution), PbI2 (99.998%), dimethyl sulfoxide (DMSO), and diethyl ether were purchased from Alfa Aesar and used without further purification. CH3NH3I was synthesized by the method described in the literature.21 Briefly, CH3NH3I was synthesized by reacting aqueous HI with CH3NH2 at 0 °C for 2 h in a three-neck flask under a N2 atmosphere with constant stirring. A white precipitate (CH3NH3I) formed during rotary evaporation of the solvent. The precipitated white powder was collected, washed three times with diethyl ether, and then dried under vacuum at 60 °C overnight. This dried CH3NH3I white powder was eventually stored in a glovebox for further fabrication. The ZnO sol−gel solution was synthesized according to a modified procedure reported in the literature.47 Typically, zinc acetate dihydrate, aluminum chloride hexahydrate, 2-methoxyethanol, and monoethanolamine (MEA) were used as the starting materials, solvent, and stabilizer, respectively. Zinc acetate dihydrate was first dissolved in a mixture of 2-methoxyethanol. The molar ratio of MEA to zinc acetate dihydrate was maintained at 1.0, and the concentration of zinc acetate was 0.5 M. To prepare Al-doped ZnO sol−gel solutions, the concentrations of Al as a dopant were varied (1%, 3%, and 5%) with respect to Zn, and the concentration of zinc acetate was 0.25 M. When the mixture was stirred, MEA was added dropwise. Then the resulting mixture was vigorously stirred in a cone-shaped glass beaker and settled in a water bath at 80 °C for 3 h. Finally, the sol−gel solution was aged for 1 day to obtain a clear and transparent homogeneous solution. Device Fabrication and Characterization. Substrates of the cells are fluorine-doped tin oxide (FTO) conducting glass (Ruilong; thickness 2.2 mm, sheet resistance 14 Ω/square). Before use, the FTO glass was first washed with mild detergent, rinsed with distilled water several times and subsequently with ethanol in an ultrasonic bath, and finally dried under an air stream. The ZnO/AZO bilayer thin films were sequentially deposited on the FTO glass substrate via spin-coating. The 237

DOI: 10.1021/acs.jpcc.7b11245 J. Phys. Chem. C 2018, 122, 236−244

Article

The Journal of Physical Chemistry C coating solution was dropped onto the FTO substrate which was rotated at 3000 rpm for 30 s, followed by thermal annealing at 200 °C for 30 min on a hot plate to evaporate the solvent and remove organic residuals. Next the substrates were transferred to a glovebox for further deposition of the perovskite active layer through the two-step spin-coating method. PbI2 (48 wt %)+KCl (1 wt %) additive was dissolved in 1 mL of dimethyl sulfoxide; CH3NH3I (4.25 wt %) was dissolved in 1 mL of 2-propanol solvent. Both solutions were kept on a hot plate at 70 °C overnight. A hot PbI2 solution was spin-coated onto ZnO thin film and annealed directly (70 °C, 10 min). The hot CH3NH3I solution was then spin-coated onto the PbI2 film; the structure was kept on the hot plate at 100 °C for 5 min to form a crystalline perovskite film. The holetransporting spiro-OMeTAD material (5 wt %), 28.5 μL of 4tert-butylpyridine (tBP) solution, and 17.5 μL of lithium bis(trifluoromethylsulfonyl)imide solution (520 mg in 1 mL of acetonitrile) all dissolved in 1 mL of chlorobenzene (CB) were further spin-coated on top. Finally, the device was completed through sequential thermal evaporation of MoO3 (8 nm) and a silver electrode (80 nm) through a shadow mask under vacuum (pressure: 1 × 10−6 Torr). The active area of each device was 10 mm2. Ultraviolet photoelectron spectroscopy (UPS) was performed by using a PHI 5000 Versa Probe apparatus equipped with an Al Kα X-ray source (1486.6 eV) and He (I) (21.22 eV) as monochromatic light sources. Absorption spectra of the films were measured with the use of a Jacobs V670 UV−vis spectrophotometer. The morphology of the ZnO, ZnO/AZO bilayer, and perovskite thin films were investigated by field-emission scanning electron microscopy (FEI Nova 200, 10 kV). Crytallographic information was obtained by X-ray diffraction (XRD) on a Bruker D8 X-ray diffractometer (2θ range: 10−60°; step size: 0.008°) equipped with a diffracted beam monochromator set for Cu Kα radiation (λ = 1.54056 Å). The photocurrent density−voltage (J−V) characteristics of the cells were illuminated inside a glovebox by a Xe lamp as a solar simulator (Thermal Oriel 1000 W), which provided a simulated AM 1.5 spectrum (100 mW cm−2). The light intensity was calibrated by using a monosilicon photodiode with a KG-fivecolor filter (Hamamatsu). External quantum efficiency (EQE) spectra were measured under monochromatic illumination (Enlitech, QE-R3011). Devices were encapsulated before they were removed for EQE measurement. The microphotoluminescence spectra (μ-PL, Horiba Jobin Yvon HR-800) of the ZnO, ZnO/AZO bilayer, and perovskite thin films were obtained by using a 325 nm He-Cd CW laser as the excitation source with a 2400 grooves/mm grating in the backscattering geometry. Time-resolved PL measurements were carried out by a nanosecond pulsed diode laser head with a wavelength of 485 nm (LDH-D-C-485 Picoquant, GmbH), a Picoharp 300 (Picoquant, GmbH) photon counting instrument, and a single photon avalanche photodetector (Picoquant, GmbH). All of the measurements were performed at room temperature.

Figure 1. (a) Device architecture of FTO/ZnO/AZO/CH3NH3PbI3/ spiro-OMeTAD/MoO3/Ag used in the study. (b) The corresponding cross-sectional SEM image of the FTO/ZnO/AZO/perovskite device.

band compared to that of ZnO, the high electron density and strong built-in electric field at the interface with perovskite would restrain the electron−hole recombination. The corresponding cross-sectional image of the FTO/ZnO/AZO/ perovskite device is shown in Figure 1(b). The ZnO/AZO thin layer was deposited on the FTO substrate by spin-coating. Then the CH3NH3PbI3 or perovskite active layer was deposited onto the ZnO/AZO bilayer using a two-step spin-coating method. A uniform distribution of perovskite film was formed on the top of the ZnO/AZO bilayer. The PHJ perovskite devices were finally fabricated using spiro-OMeTAD as a hole transporting layer (HTL), MoO3, and silver as top electrode. The thicknesses of the ZnO and AZO layers were varied to optimize the peformance of the solar cells (see Supporting Information, Figure S1). The best thicknesses of the ZnO and AZO layers found are ∼30 nm and ∼10 nm, respectively. To investigate the electronic structures of the ZnO/AZO bilayer and evaluate the effect of the AZO interfacial modification, UPS and absorption measurements were performed on the ZnO and AZO thin films. Energy-level diagrams of other Al-doping concentrations (3% and 5%) can be found in Figure S2. Figure 2(a) presents the UPS spectra of the ZnO and AZO thin films. The low binding energy tail (Eonset) of the UPS spectra for ZnO and AZO samples is shown in Figure 2(a). The VBM energy levels (relative to vacuum level) are calculated by using eq 1.48 −VBM = ℏν − (Ecutoff − Eonset)

(1)

where ℏν = 21.22 eV is the incident photon energy and the high binding energy cutoff (Ecutoff) spectra of ZnO and AZO obtained from Figure 2(a) are 17.06 and 17.10 eV, respectively. The corresponding VBM energy levels of ZnO and AZO are −7.18 and −7.20 eV. To determine the positions of the conduction band minimum (CBM) of ZnO and AZO, the optical bandgap energies are required. The optical bandgap energies (Eg) of ZnO and AZO were estimated using eq 2.49



(αℏν)n = A(ℏν − Eg )

RESULTS AND DISCUSSION Figure 1(a) illustrates the device architecture of PHJ perovskite solar cells with the ZnO/AZO bilayer used in this study. The device structure is FTO/ZnO/AZO/CH3NH3PbI3/spiroOMeTAD/MoO3/Ag, where the ZnO/AZO bilayer was deposited onto the FTO electrode as ETL. To efficiently extract electrons from the CH3NH3PbI3 active layer, a 10 nm thick sol−gel processed AZO layer was grown at the ZnO/ perovskite interface. Because AZO has a higher conduction

(2)

where n = 1/2 for indirect transition, n = 2 for direct transition, A is the proportional constant, α is the absorption coefficient, ℏ is Planck’s constant, and ν is the frequency of the incident photon. Here Eg was determined by using (αhν)2 = A(ℏν − Eg) because ZnO and AZO are both direct transition in nature. As shown in Figure S3, the corresponding optical bandgap of ZnO and AZO are 3.28 and 3.40 eV which were obtained by extrapolating the linear region of the curves near the onset of the absorption edge to the energy axis. From the positions of 238

DOI: 10.1021/acs.jpcc.7b11245 J. Phys. Chem. C 2018, 122, 236−244

Article

The Journal of Physical Chemistry C

Figure 2. (a) UPS spectra for the ZnO layer (red curve) and ZnO/ AZO bilayer (blue curve). Left: onset region; middle: the whole UPS photoemission spectra; right: cut-off region. (b) Energy-level diagram of the ZnO/AZO bilayer−perovskite device.

VBM and optical bandgap obtained from the absorption spectra, the estimated conduction band minimum (CBM) positions with respect to vacuum level (at 0 eV) are −3.90 and −3.80 eV for ZnO and AZO, respectively. In general, the optical bandgap broadening of AZO can be attributed to the increase in the carrier concentration which blocks the lowest states in the conduction band based on the Burstein−Moss effect.44 The corresponding energy level diagram is illustrated in Figure 2(b). The Fermi level values of each layer were aligned due to thermodynamic equilibrium that arises when they are combined. It was revealed that the conduction band minimum (CBM) of AZO is slightly higher than that of ZnO (ΔEV = 0.1 eV), thus indicating that this ZnO/AZO bilayer band alignment could benefit electron extraction from the CH3NH3PbI3 active layer, suppress charge accumulation in the ZnO layer, and restrain charge recombination at the ZnO/ perovskite interface. These results suggest that the sequencing of the ZnO/AZO bilayer structure is influential in tuning the energy level alignment and increase in voltage output of the device. The comparison of band profiles (including the bandbending effect due to doping and charge transfer) of ZnO/ CH3NH3PbI3 and ZnO/AZO/CH3NH3PbI3 interfaces with (green dashed lines) and without (blue solid lines) the effect of electron−hole interaction after photoexcitation is shown in Figure 3. Because the Fermi level (with respect to vacuum) in FTO (n-contact) is 0.2 eV higher than that of Ag (p-contact), electrons must be transferred from the n side to p side to make their Fermi levels aligned when FTO and Ag are brought into contact in the solar cell device, leaving surface charges near the ohmic contacts at both ends. This leads to a nearly constant

Figure 3. Schematic diagrams of field-effect-induced band bending in the heterojunctions of (a) ZnO/CH3NH3PbI3 and (b) ZnO/AZO/ CH3NH3PbI3 without illumination (solid blue line) and with illumination (dashed green line). “++++” in AZO region indicates the positive charges left after depletion of free electrons.

electric field in the active region. Most of the voltage drop should appear in the perovskite region, because it is intrinsic. The qualitative behavior of band bending depicted in Figure 3 has taken into account the built-in voltage and the solution to the Poisson equation. ∂ 2V (z)/∂z 2 = 4πe 2[Nd(z) − n(z)]/ε(z)

(3)

where V(z), n(z), Nd(z), and ε(z) denote the electrostatic potential energy, electron carrier density, dopant concentration, and dielectric constant at position z, respectively. In Figure 3(a), the band profile of ZnO has a weak band bending (because the built-in electric field is partially screened by the free carriers in ZnO, which is slightly n-type), thus having less capability to separate photocreated electron−hole pairs near the interface. An extra electron-trapping potential can be created (as indicated by the dotted line in Figure 3) by the electron−hole attraction after photoexcitation. These trapped electrons in the triangular well can hinder the charge separation process, and part of them will recombine with holes in CH3NH3PbI3. In Figure 3(b), a thin AZO layer (which has higher free electron density than ZnO due to Al doping) is inserted between ZnO and CH3NH3PbI3. Because the conduction band minimum (CBM) of AZO is higher than the CBM of ZnO, free electrons must be transferred from the 239

DOI: 10.1021/acs.jpcc.7b11245 J. Phys. Chem. C 2018, 122, 236−244

Article

The Journal of Physical Chemistry C

The XRD patterns of FTO/ZnO/AZO/perovskite, FTO/ ZnO/perovskite, and FTO/ZnO/AZO are displayed in Figure 5. The sol−gel-processed ZnO/AZO bilayer film shows an

AZO layer to the ZnO/FTO interface region, leaving a depleted AZO region with net positive charges of dopants, which are assumed to be uniformly distributed in the thin layer. Consequently, there is a strong band bending in the AZO region after adding the self-consistent potential which satisfies the Poisson equation (eq 3), as seen in Figure 3(b). The curvature of the band bending is proportional to the dopant concentration due to eq 3. After photoexcitation, the attractive electron−hole interaction will reduce the band bending but not enough to form a trapping potential at the AZO/perovskite interface. Thus, the charge separation process in the ZnO/ AZO/perovskite interface can be enhanced in comparison to the ZnO/perovskite interface. Due to the thinness of the inserted AZO layer, the electron mobility of the bilayer ETL should remain similar to that in the single layer ZnO ETL. The surface morphologies of the pure ZnO ETL, the ZnO/ AZO bilayer ETL, and the active layer of CH3NH3PbI3, respectively, are shown in Figure 4. As observed in Figure 4(a),

Figure 5. XRD patterns of FTO/ZnO/AZO/perovskite, FTO/ZnO/ perovskite, and FTO/ZnO/AZO.

amorphous structure with no crystallinity (2θ in the range of 10° to 60°) which is consistent with results in the literature47 for low-temperature-processed ZnO ETLs. The diffraction pattern of CH3NH3PbI3 film reveals peaks associated with the (002), (211), (202), (004), (310), (312), (224), (314), and (404) planes, respectively, confirming the formation of a tetrahedral perovskite crystal structure. The peak positions indicate that a pure perovskite phase was obtained; no secondary phases arising from incomplete formation of perovskite appeared (e.g., PbI2 or CH3NH3I). The J−V characteristics and EQE (external quantum efficiency) spectra of the cells based on ZnO and ZnO/AZO bilayer ETLs are shown in Figure 6(a) and 6(b). The photovoltaic parameters of devices with best performance are summarized in Table 1. The device employing ZnO/AZO bilayer ETLs shows significant enhancement in photovoltaic parameters with respect to that of the ZnO single layer ETL. The comparison of AZO ETLs with different Al-doping concentrations (1%, 3%, 5%) is shown in Figure S4. The device with a ZnO/AZO bilayer presents distinct improvement in Jsc (from 18.2 to 20.6 mA·cm−2) and Voc (from 1.04 to 1.09 V), resulting in the best PCE (from 12.3% to 16.1%). Notably, the fill-factor (FF) value was also largely improved from 64.7% to 71.6% by adding AZO as the buffer layer. Similar to the other planar device based on a normal structure of cp-TiO2/ perovskite/spiro-OMeTAD, the present structure also showed hysteresis in the J−V measurement (Figure S5), depending on the scanning direction. As shown in Figure S5(a), severe hysteresis occurred in the device employing a single ZnO ETL that can be ascribed to the unbalanced charge accumulation at the two interfaces51 of ZnO/perovskite/spiro-OMeTAD. However, the device comprising the ZnO/AZO bilayer ETL shows a greatly improved hysteresis. According to the previous report,24 the hysteresis is very small in the inverted PHJ perovskite solar cells due to better balance in electron and hole transport in the structure which uses high-conducting PCBM and PEDOT:PSS as an ETL and HTL. Therefore, we found that faster electron transfer between CH3NH3PbI3 and the ZnO/AZO bilayer made the charge transport more balanced, thus reducing the current hysteresis. On the basis of the slope (dJ/dV)−1 of the corresponding J−V curves at Jsc, we obtained a

Figure 4. SEM top-view images of (a) ZnO (inset is the bare FTO substrate) (the areas of visible cracks are indicated by yellow boxes), (b) ZnO/AZO, (c) perovskite film grown on ZnO, and (d) perovskite film grown on ZnO/AZO.

the morphology of ZnO single layer on FTO substrate shows a clear sol−gel processed granular feature. However, several cracks (indicated by yellow boxes) can be found in the ZnO layer which correlate with the FTO grain boundaries. These cracks could lead to direct contact between FTO and perovskite, causing serious charge recombination. On the other hand, with the AZO thin film modification, the ZnO/ AZO bilayer [Figure 4(b)] forms a smooth and homogeneous ETL surface without visible cracks. Figure 4(c,d) shows topview SEM images of perovskite films deposited on ZnO and ZnO/AZO bilayer thin films. The less-compact perovskite morphology observed in the film grown on ZnO [Figure 4(c)] may result in poorer connectivity between adjacent crystallites, leading to a more tortuous pathway for carrier transport and a greater likelihood of charge recombination.50 In contrast, the uniform and densely packed perovskite film [Figure 4(d)] with grain sizes at approximately ∼150−450 nm was formed on top of the ZnO/AZO bilayer film. 240

DOI: 10.1021/acs.jpcc.7b11245 J. Phys. Chem. C 2018, 122, 236−244

Article

The Journal of Physical Chemistry C

Figure 6. (a) J−V characteristics of perovskite solar cells incorporating ZnO and ZnO/AZO as ETLs measured under AM 1.5 G illumination (100 mW cm−2). (b) EQE spectra of the corresponding cells. (c) Absorption and steady-state PL spectra of perovskite films. (d) Time-resolved PL spectra of perovskite films comprising ZnO and ZnO/AZO ETLs.

Table 1. Photovoltaic Performance Parameters of the Best Performing Devices Prepared with Only ZnO Layer and with ZnO/ AZO Bilayer as ETLs cathode

Voc (V)

Jsca (mA/cm2)

Jscb (mA/cm2)

FF (%)

PCE (%)

RSH (Ω·cm2)

FTO/ZnO FTO/ZnO/AZO

1.04 ± 0.03 1.09 ± 0.03

18.2 ± 0.3 20.6 ± 0.3

17.6 ± 0.3 19.6 ± 0.4

64.7 ± 1 71.6 ± 1

12.3 ± 0.3 16.1 ± 0.4

394 ± 3% 21153 ± 3%

a Measured Jsc values from the AM 1.5 G solar simulator. bIntegrated Jsc values from the EQE measurement. cThe histograms for photovoltaic parameters of 50 cells are illustrated in Figure S6.

Table 2. Fitted Time-Resolved PL Parameters of Perovskite Films Prepared with Only ZnO Layer and with ZnO/AZO Bilayer as ETLs cathode

A1 (%)

τ1 (ns)

A2 (%)

τ2 (ns)

τave (ns)

ZnO ZnO/AZO

63.7 ± 0.5 72.1 ± 0.5

2.1 ± 0.1 1.5 ± 0.1

36.3 ± 0.5 27.9 ± 0.5

13.8 ± 0.1 6.2 ± 0.1

6.4 ± 0.1 2.8 ± 0.1

low shunt resistance (RSH) of 394 Ω·cm2 for the ZnO single layer film and a high RSH of 21153 Ω·cm2 for the ZnO/AZO film. This translates to a significant improvement in FF. The increased Jsc in the ZnO/AZO bilayer is supported by the enhancement in EQE [see Figure 6(b)]. It is found that the PHJ perovskite device comprising the ZnO/AZO bilayer ETLs exhibits a higher spectral response with a maxima EQE of 83.7% in contrast to a lower EQE of 80.9% for the device comprising the ZnO ETL. Moreover, the increased EQE of the device based on the ZnO/AZO bilayer ETLs is due to the higher light absorption within a range of 300 to 400 nm [see Figure 6(c)] and a better charge collection efficiency. The discrepancy (∼5%) between the integrated Jsc from EQE and Jsc obtained from the J−V curve is generally observed in the perovskite cell due to the difference in measurement methods.

To investigate the origin of enhanced Jsc, PL measurements were carried out to study the capability for electron transfer from perovskite to ETLs. As shown in Figure 6(c), strong PL quenching occurred in perovskite thin films grown on a ZnO/ AZO bilayer ETL compared to that grown on a ZnO ETL and indicates high electron transfer efficiency and exciton dissociation of the ZnO/AZO bilayer ETL configuration. Therefore, the higher PL quenching in the perovskite film based on the ZnO/AZO bilayer suggests faster deactivation of the excited state by the efficient charge transfer between perovskite and the ZnO/AZO bilayer thus giving rise to enhanced Jsc. Detailed PL quenching comparisons of perovskite films on bare glass and FTO are illustrated in Figure S7. To further confirm the faster electron transport properties in the ZnO/AZO bilayer ETL, time-resolved PL measurements were 241

DOI: 10.1021/acs.jpcc.7b11245 J. Phys. Chem. C 2018, 122, 236−244

Article

The Journal of Physical Chemistry C conducted and are shown in Figure 6(d). The PL lifetime was fitted with a biexponential decay function of the form I(t) = A1 exp(−t/τ1) + A2 exp(−t/τ2), where I(t) is the intensity at time t after the excitation pulse. By fitting the biexponential function to the fluorescence-decay curve, the lifetime components τ1 and τ2, as well as the amplitudes A1 and A2, can be deduced. The fitted results are summarized in Table 2. The fast decay (τ1) process is attributed to the recombination of free carriers near the interfaces in the perovskite where carriers can escape through transport to the ETL or HTL, whereas the slow decay (τ2) process is attributed to the radiative recombination of free carriers away from the interfaces. The τ1 of the ZnO/AZO/ perovskite/spiro-OMeTAD sample decreased to 1.5 ns, which was much smaller than that of the ZnO/perovskite/spiroOMeTAD sample (2.1 ns). This indicated that charger transfer between perovskite and the ZnO/AZO bilayer is faster than that without the AZO interlayer. The results of PL quenching and lifetime determination are consistent with more efficient charge separation due to the staircase band alignment of our device structure with the ZnO/AZO bilayer ETL. It is thus evident that the electrical properties are strongly influenced by employing AZO as the interlayer. Electrochemical impedance measurements were carried out for perovskite solar cells containing ZnO and ZnO/AZO as ETLs to investigate charge carrier transport properties systematically. Figure 7 shows Nyquist plots (imaginary versus

Figure 8. (a) Thermal stability test (with heating at 100 °C) and (b) normalized PCE as a function of time (in days) for perovskite solar cells containing ZnO and ZnO/AZO ETLs.

°C). We found that the CH3NH3PbI3 film deposited on the AZO interlayer shows better thermal stability than that deposited on the ZnO ETL. The instability of the CH3NH3PbI3 film deposited on the ZnO ETL was mainly due to the reaction of the CH3NH3+ proton in the perovskite layer with the surface hydroxyl groups and residual acetate ligands of the sol−gel processed ZnO.38 Previous reports on the thermal stability test of perovskite solar cells with only the ZnO ETL confirmed their severe instability against annealing time. Thus, the addition of the AZO interlayer improves its thermal stability, which can be ascribed to the decrease in Lewis acid−base chemical reaction between the perovskite and ZnO because the basic property of ZnO has been weakened by doping with aluminum.42,43 Moreover, we compared the durability of the unencapsulated perovskite solar cells containing pure ZnO ETLs with those containing ZnO/AZO ETLs in Figure 8(b). (All testing devices were stored inside a glovebox with 30 ppm oxygen and 0.1 ppm water for 45 days, and their performance was constantly monitored). The device containing ZnO/AZO as the ETL exhibits long-term stability with only 9% degradation of its initial PCE after being kept in the dark for over 45 days. In contrast, the stability of the device prepared with ZnO as the ETL shows deterioration in performance after 12 days of storage with 53% degradation of its initial PCE. The significant degradation probably arises from the ZnO/perovskite interface (excluding the oxidation effect of spiro-OMeTAD and silver electrodes). To further understand the origin of the decomposition of perovskite layers on different interfaces, XPS measurement was carried out to analyze the O1s core level spectra for both ZnO and AZO samples (as shown in Figure S8). The lower binding energy of OZn−O peak (531.08 eV) that is attributed to O atoms in the wurtzite structure of ZnO33 for AZO thin film shifts toward lower energy by 0.11 eV in

Figure 7. Nyquist plots of perovskite solar cells containing ZnO and ZnO/AZO ETLs under 1 sun illumination with applied bias voltage of 1 V.

real part of the impedance) of perovskite solar cells containing ZnO and ZnO/AZO as ETLs under 1 sun illumination at an applied bias voltage of V = 1 V. The impedance spectra consist of high frequency (low Z′) and low frequency (high Z′) features. In most cases, these features were represented by an arc or semicircle depending on the sample and the applied voltage. In the perovskite solar cell containing only ZnO as the ETL, a high resistance represented by a large arc was observed. Higher resistance could be attributed to the presence of cracks [see SEM image in Figure 3(a)] that allowed direct contact between FTO and perovskite, causing deleterious charge recombination that could yield to poor device performance. In contrast, for the perovskite solar cells containing ZnO/AZO as ETLs, the addition of AZO decreased the resistance (corresponding to a smaller arc in Figure 7), thus indicating a smoother carrier movement that leads to less recombination and better device performance. Figure 8 (a) shows the thermal stability test of perovskite films deposited on ZnO and ZnO/AZO ETLs (heating at 100 242

DOI: 10.1021/acs.jpcc.7b11245 J. Phys. Chem. C 2018, 122, 236−244

Article

The Journal of Physical Chemistry C comparison with ZnO thin film. Moreover, the intensity of OZn−O peak in the AZO thin film is lower than that in the ZnO thin film which can be explained by Al doping.42 The higher binding energy of the second peak (532.97 eV), corresponding to the OH− or O−2 ions in AZO, shifts toward higher binding energy and has lower intensity than in the ZnO thin film. The above result indicates that the AZO thin film is more oxygen deficient. Therefore, OH− group passivation enhances the thermal stability and durability of perovskite solar cells containing the ZnO/AZO ETL.

thanks the Career Development Award of Academia Sinica, Taiwan (103-CDA-M01) for financial support. We thank Mr. Chi-Ching Liu for the time-resolved PL measurements.



(1) De Wolf, S.; Holovsky, J.; Moon, S. J.; Loper, P.; Niesen, B.; Ledinsky, M.; Haug, F. J.; Yum, J. H.; Ballif, C. Organometal halide perovskites: Sharp optical absorption edge and its relation to photovoltaic performance. J. Phys. Chem. Lett. 2014, 5, 1035−1039. (2) Kim, H. S.; Lee, C. R.; Im, J. H.; Lee, K. B.; Moehl, T.; Marchioro, A.; Moon, S. J.; Humphry-Baker, R.; Yum, J. H.; Moser, J. E.; Gratzel, M.; Park, N.-G. Lead iodide perovskite sensitized all-solidstate submicron thin film mesoscopic solar cell with efficiency exceeding 9%. Sci. Rep. 2012, 2, 591. (3) Stranks, S. D.; Eperon, G. E.; Grancini, G.; Menelaou, C.; Alcocer, M. J.; Leijtens, T.; Herz, L. M.; Petrozza, A.; Snaith, H. J. Electron-hole diffusion lengths exceeding 1 micrometer in an organometal trihalide perovskite absorber. Science 2013, 342, 341− 344. (4) Wehrenfennig, C.; Eperon, G. E.; Johnston, M. B.; Snaith, H. J.; Herz, L. M. High charge carrier mobilities and lifetimes in organolead trihalide perovskites. Adv. Mater. 2014, 26, 1584−1589. (5) You, J.; Hong, Z.; Yang, Y. M.; Chen, Q.; Cai, M.; Song, T. B.; Chen, C. C.; Lu, S.; Liu, Y.; Zhou, H.; Yang, Y. Low-temperature solution-processed perovskite soalr cells with high efficiency and flexibility. ACS Nano 2014, 8, 1674−1680. (6) Wojciechowski, K.; Saliba, M.; Leijtens, T.; Abate, A.; Snaith, H. J. Sub-150°C processed meso-superstructured perovskite solar cells with enhanced efficiency. Energy Environ. Sci. 2014, 7, 1142−1147. (7) Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal halide perovskites as visible-light sensitizers for photovoltaic solar cells. J. Am. Chem. Soc. 2009, 131, 6050−6051. (8) Yang, W. S.; Park, B. W.; Jung, E. H.; Jeon, N. J.; Kim, Y. C.; Lee, D. U.; Shin, S. S.; Seo, J.; Kim, E. K.; Noh, J. H.; Seok, S. Iodide management in formamidinium-lead-halide-based perovskite layers for efficient solar cells. Science 2017, 356, 1376−1379. (9) Zhao, Y.; Zhu, K. Charge transport and recombination in perovskite in (CH3NH3)PbI3 sensitized TiO2 solar cells. J. Phys. Chem. Lett. 2013, 4, 2880−2884. (10) Lee, M. M.; Teuscher, J.; Miyasaka, T.; Murakami, T. N.; Snaith, H. J. Efficient hybrid solar cells based on meso-structured organometal halide perovskites. Science 2012, 338, 643. (11) Yang, W. S.; Noh, J. H.; Jeon, N. J.; Kim, Y. C.; Ryu, S.; Seo, J.; Seok, S. I. High-performance photovoltaic perovskite layers fabricated through intramolecular exchange. Science 2015, 348, 1234. (12) Burschka, J.; Pellet, N.; Moon, S. J.; Humphry-Baker, R.; Gao, P.; Nazeeruddin, M. K.; Gratzel, M. Sequential deposition as a route to high-performance perovskite-sensitized solar cells. Nature 2013, 499, 316. (13) Eperon, G. E.; Burlakov, V. M.; Docampo, P.; Goriely, A.; Snaith, H. J. Morphological control for high performance, solutionprocessed planar heterojunction perovskite solar cells. Adv. Funct. Mater. 2014, 24, 151−157. (14) Liu, M. Z.; Johnston, M. B.; Snaith, H. J. Efficient planar heterojunction perovskite solar cells by vapour and deposition. Nature 2013, 501, 395. (15) Ren, Y. K.; Ding, X. H.; Wu, Y. H.; Zhu, J.; Hayat, T.; Alsaedi, A.; Xu, Y. F.; Li, Z. Q.; Yang, S. F.; Dai, S. Y. Temperature-assisted rapid nucleation: a facile method to optimize the film morpgology for perovskite solar cells. J. Mater. Chem. A 2017, 5, 20327−20333. (16) Ball, J. M.; Lee, M. M.; Hey, A.; Snaith, H. J. Low-temperature processed meso-superstructured to thin-film perovskite solar cells. Energy Environ. Sci. 2013, 6, 1739. (17) Wang, J. T. W.; Ball, J. M.; Barea, E. M.; Abate, A.; Alexander Webber, J. A.; Huang, J.; Saliba, M.; Mora-Sero, I.; Bisquert, J.; Snaith, H. J.; Nicholas, R. J. Low-temperature process electron collection layers of graphene/TiO2 nanocomposite in thin film perovskite solar cells. Nano Lett. 2014, 14, 724−730.



CONCLUSION In summary, we have successfully demonstrated how the lowtemperature solution-processed ZnO/AZO bilayer ETL improves the efficiency of PHJ perovskite solar cells. In this study, the PCE obtained can exceed 16%. The introduction of an AZO layer not only helps the formation of uniform surface morphology to fill the defects but also leads to a staircase band alignment in the ZnO/AZO/perovskite heterostruucture, which is shown to yield better electron extraction, thereby significantly enhancing device performance. It is noteworthy that the results obtained for our optimized device surpass the state-of-the-art PCE and Voc of solution-processed ZnO-based perovskite solar cells without any polymer interlayer. Moreover, the device prepared with ZnO/AZO bilayer ETLs showed longterm stability: only 9% degradation of its initial PCE after being kept in the dark for over 45 days. This approach paves a new way for the development of low-cost perovskite solar cells.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcc.7b11245. J−V curves at different molar concentrations of ZnO and ZnO/AZO ETLs, UPS spectra of 3% AZO and 5% AZO films and the corresponding energy-level diagram, Tauc plots for optical bandgap extraction, J−V curves at different concentrations of AZO interlayers, J−V hysterisis characteristics, histograms of device performance parameters, steady-state PL spectra of perovskite films on different interfaces, and XPS spectra of O1s of ZnO and AZO 1% films (PDF)



REFERENCES

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. ORCID

Chih-Wei Chu: 0000-0003-0979-1729 Yia-Chung Chang: 0000-0003-1851-4651 Author Contributions

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported in part by the Ministry of Science and Technology of Taiwan under contract nos. MOST 104-2112M-001-009-MY2 and 104-2221-E-001-014-MY3. C. W. Chu 243

DOI: 10.1021/acs.jpcc.7b11245 J. Phys. Chem. C 2018, 122, 236−244

Article

The Journal of Physical Chemistry C (18) Etgar, L.; Gao, P.; Xue, Z.; Peng, Q.; Chandiran, A. K.; Liu, B.; Nazeeruddin, M. K.; Gratzel, M. Mesoscopic CH3NH3PbI3/TiO2 heterojunction solar cells. J. Am. Chem. Soc. 2012, 134, 17396−17399. (19) Chen, Q.; Zhou, H.; Hong, Z.; Luo, S.; Duan, H. S.; Wang, H. H.; Liu, Y.; Li, G.; Yang, Y. Planar heterojunction perovskite solar cells via vapor-assisted solution process. J. Am. Chem. Soc. 2014, 136, 622− 625. (20) Chen, C. W.; Kang, H. W.; Hsiao, S. Y.; Yang, P. F.; Chiang, K. M.; Lin, H. W. Efficient and uniform planar-type perovskite solar cells by simple sequential vacuum deposition. Adv. Mater. 2014, 26, 6647− 6652. (21) Boopathi, K. M.; Mohan, R.; Huang, T. H.; Budiawan, W.; Lin, M. Y.; Lee, C. H.; Ho, K. C.; Chu, C. W. Synergistic improvements in stability and performance of lead iodide perovskite solar cells incorporating salt additives. J. Mater. Chem. A 2016, 4, 1591−1597. (22) Jeng, J. H.; Chiang, Y. F.; Lee, M. H.; Peng, S. R.; Guo, T. F.; Chen, P.; Wen, T. C. CH3NH3PbI3 perovskite/fullerence planarheterojunction hybrid solar cells. Adv. Mater. 2013, 25, 3727−3732. (23) Sun, S.; Salim, T.; Mathews, N.; Duchamp, M.; Boothroyd, C.; Xing, G.; Sum, T. C.; Lam, Y. M. The origin of high efficiency in lowtemperature solution-processed bilayer organometal halide hybrid solar cells. Energy Environ. Sci. 2014, 7, 399−407. (24) Kim, H. S.; Jang, I. H.; Ahn, N.; Guerrero, A.; Choi, M.; Bisquert, J.; Park, N. G. Control of I−V hysteresis in CH3NH3PbI3 perovskite solar cell. J. Phys. Chem. Lett. 2015, 6, 4633−4639. (25) Snaith, H. J. Perovskites: The emergence of a new era for lowcost, high efficiency solar cells. J. Phys. Chem. Lett. 2013, 4, 3623− 3630. (26) Beiley, Z. M.; McGehee, M. D. Modeling low cost hybrid tandem photovoltaics with the potential for efficiencies exceeding 20%. Energy Environ. Sci. 2012, 5, 9173−9179. (27) Ozgur, U.; Alivov, Y. I.; Liu, C.; Teke, A.; Reshchikov, M. A.; Dogan, S.; Avrutin, V.; Cho, S. J.; Morkoc, H. A comprehensive review of ZnO materials and devices. J. Appl. Phys. 2005, 98, 041301. (28) Wu, S. H.; Chan, C. H.; Chien, C. H.; Yaseen, M. T.; Liang, C. T.; Chang, Y. C. Enhanced emission and photoconductivity due to photo-induced charge transfer from Au nanoislands to ZnO. Appl. Phys. Lett. 2016, 108, 041104. (29) Wang, Z. L. Nanostructures of zinc oxide. Mater. Today 2004, 7, 26−33. (30) Lin, H. W.; Kang, H. W.; Huang, Z. Y.; Chen, C. W.; Chen, Y. H.; Lin, L. Y.; Lin, F.; Wong, K. T. An effective bilayer cathode buffer for highly efficient small molecule organic solar cells. Org. Electron. 2012, 13, 1925−1929. (31) Hsiao, S. Y.; Lin, H. L.; Lee, W. H.; Tsai, W. L.; Chiang, K. M.; Liao, W. Y.; Ren-Wu, C. Z.; Chen, C. Y.; Lin, H. W. Efficient allvacuum deposited perovskite solar cells by controlling reagent partial pressure in high vacuum. Adv. Mater. 2016, 28, 7013−7019. (32) Sekine, N.; Chou, C. H.; Kwan, W. L.; Yang, Y. ZnO nano-ridge structure and its application in inverted polymer solar cell. Org. Electron. 2009, 10, 1473−1477. (33) Ambade, R. B.; Ambade, S. B.; Mane, R. S.; Lee, S. H. Interfacial engineering importance bilayered ZnO cathode buffer on the photovoltaic performance of inverted organic solar cells. ACS Appl. Mater. Interfaces 2015, 7, 7951−7960. (34) Dong, J.; Zhao, Y.; Shi, J.; Wei, H.; Xiao, J.; Xu, X.; Luo, J.; Xu, J.; Li, D.; Luo, Y.; Meng, Q. Impressive enhancement in the cell performance of ZnO nanorod-based perovskite solar cells with Aldoped ZnO interfacial modification. Chem. Commun. 2014, 50, 13381−13384. (35) Liang, L.; Huang, Z.; Cai, L.; Chen, W.; Wang, B.; Chen, K.; Bai, H.; Tian, Q.; Fan, B. Magnetron sputtered zinc oxide nanorods as thickness-insensitive cathode interlayer for perovskite planar-heterojunction solar cells. ACS Appl. Mater. Interfaces 2014, 6, 20585−20589. (36) Mahmood, K.; Swain, B. S.; Amassian, A. 16.1% efficient hysteresis-free mesostructured perovskite solar cells based on synergysticaly improved ZnO nanorod arrays. Adv. Energy Mater. 2015, 5, 1500568.

(37) Cheng, Y.; Yang, Q. D.; Xiao, J.; Xue, Q.; Li, H. W.; Guan, Z.; Yip, H. L.; Tsang, S. W. Decomposition of organometal halide perovskite films on zinc oxide nanoparticles. ACS Appl. Mater. Interfaces 2015, 7 (36), 19986−19993. (38) Yang, J.; Siempelkamp, B. D.; Mosconi, E.; De Angelis, F.; Kelly, T. L. Origin of the thermal instability in CH3NH3PbI3 thin films deposited on ZnO. Chem. Mater. 2015, 27 (12), 4229−4236. (39) Kim, J.; Kim, G.; Kim, T. K.; Kwon, S.; Back, H.; Lee, J.; Lee, S. H.; Kang, H.; Lee, K. Efficient planar-heterojunction perovskite solar cells achieved via interfacial modification of sol-gel ZnO electron collection layer. J. Mater. Chem. A 2014, 2, 17291−17296. (40) Qiu, W. M.; Buffiere, M.; Brammertz, G.; Paetzold, U. W.; Froyen, L.; Heremans, P.; Cheyns, D. High efficiency perovskite solar cells using a PCBM/ZnO double electron transport layer and a short air-aging step. Org. Electron. 2015, 26, 30−35. (41) An, Q.; Fassl, P.; Hofstetter, Y. J.; Becker-Koch, D.; Bausch, A.; Hopkinson, P. E.; Vaynzof, Y. High performance planar perovskite solar cells by ZnO transport layer engineering. Nano Energy 2017, 39, 400−408. (42) Zhao, X. Y.; Shen, H. P.; Zhang, Y.; Li, X.; Zhao, X. C.; Tai, M. Q.; Li, J. F.; Li, J. B.; Li, X.; Lin, H. Aluminum-doped zinc oxide as highly stable elctron collection layer for perovskite solar cells. ACS Appl. Mater. Interfaces 2016, 8, 7826−7833. (43) Tseng, Z. L.; Chiang, C. H.; Chang, S. H.; Wu, C. G. Surface engineering of ZnO electron transporting layer via Al-doping for high efficiency planar perovskite solar cells. Nano Energy 2016, 28, 311− 318. (44) Lin, J. P.; Wu, J. M. The effect of annealing process on electronic properties of sol-gel derived Al-doped ZnO films. Appl. Phys. Lett. 2008, 92, 134103. (45) Tan, H.; et al. Perovskites: The emergence of a new era for lowcost, high efficiency solar cells. Science 2017, 355, 722−726. (46) Liu, D.; Kelly, T. L. Perovskite solar cells with a planar heterojunction structure prepared using room-temperature solution processing techniques. Nat. Photonics 2014, 8, 133−138. (47) Sun, Y.; Seo, J. H.; Takacs, C. J.; Seifter, J.; Heeger, A. J. Inverted polymer solar cells integrated with a low-temperatureannealed sol-gel-derived ZnO film as an electron transport layer. Adv. Mater. 2011, 23, 1679−1683. (48) Hu, T.; Li, F.; Yuan, K.; Chen, Y. Efficiency and air-stability improvement of flexible inverted polymer solar cells using ZnO/ Poly(ethylene glycol) hybrids as cathode buffer layers. ACS Appl. Mater. Interfaces 2013, 5, 5763−5770. (49) Mane, R. S.; Lee, W. J.; Pathan, H. M.; Han, S. H. Nanocrystalline TiO2/ZnO thin films: Fabrication and application to dye-sensitized solar cells. J. Phys. Chem. B 2005, 109, 24254−24259. (50) Liu, D.; Gangishetty, M. K.; Kelly, T. L. Effect of CH3NH3PbI3 thickness on device efficiency in planar heterojunction perovskite solar cells. J. Mater. Chem. A 2014, 2, 19873−19881. (51) Chen, W.; Wu, Y.; Yue, Y.; Liu, J.; Zhang, W.; Yang, X.; Chen, H.; Bi, E.; Ashraful, I.; Gratzel, M.; Han, L. Efficient and stable largearea perovskite solar cells with inorganic charge extraction layers. Science 2015, 350, 944−948.

244

DOI: 10.1021/acs.jpcc.7b11245 J. Phys. Chem. C 2018, 122, 236−244